You are on page 1of 12

Geophysical Journal International

Geophys. J. Int. (2011) 186, 825836

doi: 10.1111/j.1365-246X.2011.05073.x

On standard and optimal designs of industrial-scale 2-D seismic


surveys
T. Guest1,2 and A. Curtis1,2
1 School

of GeoSciences, The University of Edinburgh, Grant Institute, The Kings Buildings, West Mains Road, Edinburgh, EH9 3JW, UK.
E-mail: t.e.guest@sms.ed.ac.uk
2 ECOSSE (Edinburgh Collaborative of Subsurface Science and Engineering)

Accepted 2011 May 5. Received 2011 May 5; in original form 2010 July 23

1 I N T RO D U C T I O N
Large sums of money are invested every year in geophysical surveys
and experiments by both academia, governmental organizations and
industry to constrain physical properties of the Earths subsurface.
Before any data is collected a survey design process must be performed, the aim of which is to maximize the amount of target
information we expect to record whilst also taking into account any
physical, logistical and cost constraints that define bounds on the
types of experiments that are feasible. Maximizing the amount of
information we expect to record often trades off with minimizing
the cost of the survey. For this reason, optimizing the design of the
survey in terms of cost, logistics and the information the survey is
expected to provide becomes of critical importance to maximizing
return on investment (Maurer & Boerner 1998a; Curtis & Maurer
2000).
Statistical experimental design (SED), a mature field of statistics,
is focused on the development of methods to design experiments (or
surveys) so as to maximize information, typically by minimizing
the expected post-experimental uncertainties on parameters of interest whilst satisfying other necessary constraints. Although SED
is an established methodology in other scientific fields, the majority
of designs of experiments in the geosciences are based on heuris
C

2011 The Authors


C 2011 RAS
Geophysical Journal International 

tics (rules of thumb). Within geophysics, where enormous sums of


money are spent on data collection, formal SED theory has only
been applied in a limited number of cases: to design tomographic
surveys (Barth & Wunsch 1990; Curtis 1999a,b; Curtis et al. 2004;
Ajo-Franklin 2009), earthquake monitoring surveys (Kijko 1977a,b;
Rabinowitz & Steinberg 2000; Steinberg et al. 1995; Winterfors &
Curtis 2008, 2010), microseismic monitoring surveys (Curtis et al.
2004; Coles & Curtis 2011), resistivity surveys (Maurer et al. 2000;
Stummer et al. 2004; Furman et al. 2004, 2007; Wilkinson et al.
2006; Coles & Morgan 2009), electromagnetic surveys (Maurer &
Boerner 1998b), anisotropic surveys (Coles & Curtis 2010), geological expert elicitation or interrogation methods (Curtis & Wood
2004), and amplitude versus offset seismic experiments (Van den
Berg et al. 2003, 2005; Guest & Curtis 2009, 2010). While this
may seem like a significant body of literature, most of these studies
are purely synthetic: amongst them, the number of published experiments actually acquired using SED-based designs seems to be
around five.
Optimal experimental design requires an understanding of how
the recorded data are related to post-experimental parameter uncertainties (Box & Lucas 1959; Atkinson & Donev 1992). Let function
F represent the relationship between the parameters m of interest
and data d that can be recorded, such that if measurement error is

825

GJI Seismology

Key words: Inverse theory; Probability distributions; Statistical seismology.

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

SUMMARY
The principal aim of performing a survey or experiment is to maximize the desired information
within a data set by minimizing the post-survey uncertainty on the ranges of the model parameter values. Using Bayesian, non-linear, statistical experimental design (SED) methods we
show how industrial scale amplitude variations with offset (AVO) surveys can be constructed
to maximize the information content contained in AVO crossplots, the principal source of
petrophysical information from seismic surveys. The design method allows offset dependent
errors, previously not allowed in non-linear geoscientific SED methods. The method is applied to a single common-midpoint gather. The results show that the optimal design is highly
dependent on the ranges of the model parameter values when a low number of receivers is
being used, but that a single optimal design exists for the complete range of parameters once
the number of receivers is increased above a threshold value. However, when acquisition and
processing costs are considered we find that a design with constant spatial receiver separation
survey becomes close to optimal. This explains why regularly-spaced, 2-D seismic surveys
have performed so well historically, not only from the point of view of noise attenuation and
imaging in which homogeneous data coverage confers distinct advantages, but also to provide
data to constrain subsurface petrophysical information.

826

T. Guest and A. Curtis

ignored for now then data


d = F (m)

(1)

2 METHOD
We first define precisely the type of parameters of interest in our
study, the data type with which their values are to be estimated, and
the so-called forward function F relating the parameters (m) and
data (d). We then specify how the amount of available information
about the parameters can be measured or quantified.
2.1 The amplitude versus offset (AVO) crossplot
The amplitude of a seismic wave initially of unit amplitude which
is reflected from a subsurface boundary between two geological
layers at depth is a function of the incident angle of the wave at
the boundary, the density i and the elastic media properties summarized by the P-wave velocity i , and S-wave velocity i , for an
isotropic medium of layers i = 1, 2 above and below the boundary,
respectively. The recorded amplitudes of the reflected (and transmitted) waves (after accounting for geometrical spreading effects
during propagation) are given by the solution to the non-linear,
simultaneous Zoeppritz equations (Zoeppritz 1919).
Castagna & Swan (1997) introduced the notion of AVO crossplotting, where an estimate of the normal incidence P-wave reflection
coefficient (the so-called AVO intercept) is plotted against a measure of the offset-dependent reflectivity (the AVO gradient). In the
majority of cases the AVO intercept A, and gradient B, are calculated using the Shuey (1985) two-term approximation, which is
valid up to incident angles of approximately 30 ,
R () A + B sin2 ,

(2)

where R is the P-wave reflection coefficient, A is the AVO intercept,


B is the AVO gradient and is the angle of incidence. Furthermore
Shuey (1985) shows that A and B are approximately



1 
+
A
2 

(3)

 



1 
 2
2
+
,
2
B
2 



where  is the change in P-wave velocity across the interface
( 2 1 ),  is the average P-wave velocity (( 2 + 1 )/2), and
the other parameters are defined similarly. Once the intercept and
gradient have been calculated, an AVO inversion technique can
therefore be used to estimate constraints on , and for both
layers from the information contained in the AVO crossplot using
eq. (3).
In reality the intercept and gradient are estimated from discrete,
noisy data recorded at receiver locations at varying offsets from

C 2011 The Authors, GJI, 186, 825836
C 2011 RAS
Geophysical Journal International 

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

would be recorded if parameter values m were correct in the sense


that they accurately represent the true Earth. The subscript in
the function F indicates that the parameterdata relationship is
dependent on the experimental design , where is a vector representing, for example, source and receiver types and locations,
but also potentially defining different possible data processing
methods.
The principle reason that SED methods have not gained general
acceptance in the geosciences (other than a lack of awareness) is
that most research effort on SED in the statistical community has
focussed on developing methods that assume a linear or linearized
relationship F between parameters and data, while in geophysical
applications the parameterdata relationship is commonly significantly non-linear. As a result, linearized SED methods are not
necessarily robust in geophysical applications, and those fully nonlinear methods that do exist are considered to be too computationally
costly given typical numbers of parameters and data. As a result,
out of all of the above geophysical references, only the work of Van
den Berg et al. (2003, 2005), Winterfors & Curtis (2008, 2010),
Guest & Curtis (2009, 2010) and Coles & Curtis (2011) apply fully
non-linear design theory to geophysical surveys.
Guest & Curtis (2009) introduced a novel method whereby an
optimal survey design is produced through an iterative process.
First, for a given, uncertain subsurface structure defined by Pwave and S-wave velocity and density values and likelihood distributions, the single source receiver offset that provides the maximum information about the subsurface via the Zoeppritz equations
(Zoeppritz 1919) is found. Given that this receiver is now located, the
iterative method locates the next receiver such that the additional
or marginal information expected to be recorded is maximized.
This process is repeated iteratively, optimally locating further receivers until some maximum cost or minimum expected information threshold is reached. The main drawback of using the Guest &
Curtis (2009) method to design an industrial scale pre-acquisition
survey is that even though the design space remains 1-D in each
iteration (only one receiver is located at a time), each receiver also
contributes a separate dimension in data space. To locate the optimal location for the fifth receiver therefore requires the integral
of a 5-D space be evaluated for every possible receiver location in
the design space; consequently the problem becomes computationally intractable when more than about 10 sources or receivers are
required.
Guest & Curtis (2010) used the method of Guest & Curtis (2009)
to create post-acquisition processing designs (based on petrophysical knowledge) by averaging and upscaling the 10-receiver designs
to represent spatial receiver density. From a full seismic survey
dataset, Guest & Curtis (2010) were able to select which receivers
recorded the maximum information about a given subsurface layer.
Hence, although the methods of Guest & Curtis (2009, 2010) could
be applied to design pre-acquisition surveys, the main focus of work
to-date has been on data selection.
Our approach here is similar to that used by Ajo-Franklin (2009)
for linearized methods, were the design space is reparameterized
with a low number of hyperparameters that control the receiver
density, rather than individual receiver positions. By using this technique of hyperparameterization we show for the first time that the
problem of integrating fully non-linear SED design methods into
industrial-scale geophysical standard-practice is computationally
tractable.

The method presented is used to optimize the information contained within AVO crossplots directly, bringing the acquisition design stage much closer to the standard seismic processing flow
than in previous work (van den Berg et al. 2003, 2005; Guest &
Curtis 2009) in which a full inversion of the Zoeppritz equations
was assumed. We also show how to integrate variable data errors in
the design stage which has not before been included in non-linear
Geoscientific design problems. We illustrate how the methods can
be used to optimize a survey design for a common midpoint gather,
and how varying the prior subsurface knowledge and number or
receivers alters the optimal design. Finally we analyse the information gain from the optimized survey design over industry-standard
designs. The results explain why standard industrial designs have
been so successful in constraining subsurface petrophysical information.

Design of industrial-scale 2-D seismic surveys


the source. The location of the receivers can therefore affect the
accuracy to which parameters A and B can be estimated, and consequently where the reflection coefficient data plots in the AVO
crossplot. In turn, the accuracy of estimating the subsurface velocities and densities from the crossplot values are similarly affected.
In this paper, we apply non-linear SED methods to show how
the information represented in the AVO crossplot method can be
maximized for any subsurface parameters and geological model, so
as to minimize velocity and density uncertainty after applying AVO
inversion techniques. In our definitions herein, the data d represent
the AVO intercept and gradient calculated from reflection coefficient
values, given a reservoir model m and receiver density profile . d
and m are related through eq. (3).
2.2 Measuring information

(d|m, ) (m)
,
(m|d, ) =
(d| )

(4)

where (d|m, ) represents a pdf of the data d that would be observed given true parameter values m and survey design , (m)
is a pdf representing the prior information on parameters m, and
(d| ) is the marginal distribution over observed data and contains
all information about which data are likely to be recorded during
survey (Tarantola 2005).
The optimal receiver density profile corresponds to the design
that maximizes the information expected to be contained in the
posterior parameter pdf in eq. (4). Shewry & Wynn (1987) showed
that a suitable information measure
( ) can be defined as

(5)

( ) = Ent { (d| )} Ent { (d|m, )} (m) dm,


where Ent is the entropy function defined by Shannon (1948) and
represents a measure of the uncertainty represented by a pdf. Shewry

& Wynn (1987) showed that eq. (5) represents a measure of the
parameter information expected to be gained by performing the experiment. The design measure combines the uncertainty embodied
in the marginal distribution (d| ) which represents the probability distribution of the data d (AVO intercept and gradient values)
given a specific survey design (the first term on the right), and the
average data uncertainty Ent { (d|m, )} over all possible models
given the same specific survey design (second term on the right).
In cases where the data error is not design dependent, this second
integral term in eq. (5) can be assumed constant. See Guest & Curtis
(2009, 2010) for a more complete mathematical development in a
Geophysical context.
Essentially Shewry & Wynn (1987) showed that the prior data
space uncertainty as defined by eq. (5) is directly related to the
expected, post-experimental model space information: maximizing
the former with respect to design also maximizes the latter. Most
importantly though, to calculate
( ) in eq. (5) only requires that the
prior information on parameters (m) is projected through the physical relationship F (m) [to calculate (d|m, ) and (d| )]. Maximizing
( ) thus only requires that the forward function (rather
than the inverse problem) be evaluated, and doing so implies automatically that inverted model parameter uncertainties are expected
to be minimized.
2.3 AVO design method
To calculate the optimality of a specific experimental or survey
design using eq. (5) we first construct an AVO crossplot based on
prior information about the reservoir model described by (m)
and on the survey design . Whereas Guest & Curtis (2009, 2010)
assumed a constant error with increasing offset and hence assumed
that the integral term in eq. (5) was constant and thus irrelevant from
the perspective of survey design, we now consider offset-dependent
errors and therefore also include the integral term. Fig. 1(a) shows
the standard deviation of the offset-dependent Gaussian error of the
reflection coefficient that is used here, but any other such curve
could be employed in our design method.
The work of Guest & Curtis (2009, 2010) was limited by the
design space dimensionality since the location of each selected
receiver represented an additional dimension in both and d, and
due to the required entropy calculation in eq. (5) the method suffered
strongly from the curse of dimensionality (Curtis & Lomax 2001).
In this paper, we instead define the design vector to describe the

Figure 1. (a) Offset dependent reflection coefficient error. The error value represents the standard deviation of a Gaussian error. (b) Two-term Shuey equation
solution (eq. (2) - solid line) calculated from simulated data (dots) for a specific survey design and reservoir model.

C

2011 The Authors, GJI, 186, 825836


C 2011 RAS
Geophysical Journal International 

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

We adopt a Bayesian approach for parameter inference in which


probability density functions (pdfs) represent states of information about parameters, and expected post-experimental uncertainties can be quantified to assess design quality without requiring
any linearization of the forward function F (m). The terms experiment and survey used subsequently are synonymous. According
to Bayes theorem the posterior or post-experimental pdf describing information about the parameters m, given recorded data d and
survey design , is given by

827

828

T. Guest and A. Curtis

Figure 2. Receiver density profile (solid line) defined by parameters P, the


angular density at zero offset and Q, the angular density at maximum offset.
The area of the shaded section is equal to the total number of receivers.

angular density of receiver locations. In the first place we do this


using only two parameters,
(6)

where P is the angular density of receivers at vertical incidence, and


Q is the angular density at the maximum allowed incident angle;
the density values at intermediate offsets are linearly interpolated
(Fig. 2 ). This formulation of the design problem is termed hyperparameterization in some fields and reduced parameterization by
Ajo-Franklin (2009) who used it for optimizing cross-borehole tomography surveys using linearized methods. The approach does not
allow each individual receiver to be placed at an arbitrary location;
instead the optimal design is found via a set of hyperparameters,
in this case receiver density. The area under the receiver density
plot (shaded area in Fig. 2) represents the total number of receivers
placed. For a given fixed maximum incident angle I (30 in Fig. 2)
and total number of receivers N, the two design parameters are
related by
2N
P.
(7)
I
Consider the case where a total of N = 300 receivers are to be
placed over a 30 offset range (the approximate range of valid angles
of the two-term Shuey eq. 2). Note that this is almost two orders of
magnitude more receivers then have been designed previously using
non-linear design methods, and also that our approach allows almost
any number of receivers to be located with approximate optimality.
Q=


(m) Ent { (d|m, )} dm

M
1 
Ent { (d|mi , )} ,
M i=1

(8)

where M is the total number of reservoir models sampled from the


prior parameter distribution (m).
The marginal distribution (d| ) in eq. (5) is represented by
the normalized AVO crossplot histogram resulting from all of the
data realizations for all model parameter realizations (i.e. for a
representative sample of all possible data that could be collected
in the survey given the prior information on the possible range of
reservoir models).
For each survey design the expected information gain is then
calculated using eq. (5). The density profile that corresponds to
the maximum value is the optimal survey design, since that design

Figure 3. (a) Receiver density profiles for three survey designs, and (b) the corresponding cumulative number of placed receivers as a function of incident
angle. Dashed line: P = 20, Q = 0. Dotted line: P = Q = 10. Solid line: P = 0, Q = 20.

C 2011 The Authors, GJI, 186, 825836
C 2011 RAS
Geophysical Journal International 

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

= [P, Q] ,

The extreme design parameter ranges considered here are P = 0,


Q = 20 and P = 20, Q = 0, both shown in Fig. 3(a). Fig. 3(b) shows
the cumulative number of receivers placed as a function of incident
angle for the three example density profiles in Fig. 3(a). It should
also be stressed that a constant density with respect to incident
angle at the subsurface interface does not equate to constant receiver
separation in spatial receiver locations on the ground surface.
For a given reservoir model, the reflection coefficients at each of
the placed receiver locations can be calculated by solving the Zoeppritz equations. For each receiver a Gaussian error is added to the
reflection coefficient, with standard deviation shown in Fig. 1(a).
According to standard practice the two-term Shuey equation
(eq. 2) is then fit in a least-squares sense to the resulting reflection coefficient data to determine the AVO gradient B and intercept
A (eq. 3). Fig. 1(b) shows the 2-term Shuey equation solution calculated for one example of simulated data. This, however, only
constitutes a single realization of the data for one specific reservoir
model.
To accurately estimate the pdf (d|m, ), multiple realizations
of the noisy data for the same reservoir model and receiver density
distribution are required, and for each realization a separate AVO
intercept and gradient are calculated and histogrammed in a discretised AVO crossplot. The resulting crossplot represents an estimate
of the uncertainty in calculating the intercept and gradient due to
the measurement noise for the given receiver density profile. The
histogram is normalized to have unit volume whereafter it represents a numerical approximation to the pdf (d|m, ). A numerical
approximation can therefore be calculated for the integral term in
eq. (5) as

Design of industrial-scale 2-D seismic surveys


is expected to record data that will provide maximum information
about, and hence most tightly constrain, the subsurface reservoir
parameters.

Table 2. Fluid parameters required for the Goldberg &


Gurevich (1998) model. The ranges represent the extreme
values of the uniform prior pdfs used to create velocity
and density models. Values are taken from Clark (1992),
Carcione et al. (2003), Chen & Dickens (2009).

3 CMP EXAMPLE

Parameter

Range

To illustrate the use of this novel method, we apply the non-linear


design algorithm to a single CMP gather and assess which of the
prior model parameters has the largest effect on the final design
and should therefore be constrained as tightly as possible before the
survey is conducted. Since this is the first time that a truly industrialscale seismic CMP gather (potentially hundreds of sourcereceiver
pairs) is being designed using non-linear methods, we also assess
how the number of receivers used in the survey affects the receiver
distribution in the final optimal design.

Brine bulk modulus (GPa)


Oil bulk modulus (GPa)
Gas bulk modulus (GPa)
Brine density (kg m3 )
Oil density (kg m3 )
Gas density (kg m3 )

2.43.2
0.50.75
0.01
10401090
616738
100

3.1 Reservoir model

Table 1. Rock parameters required for the Goldberg &


Gurevich (1998) model. The ranges represent the extreme
values of the uniform prior pdfs used to create velocity
and density models. Extreme values are taken from Marion et al. (1992), Mavko et al. (1998), Carcione et al.
(2003), Chen & Dickens (2009).


C

Parameter

Range

Sand bulk modulus (GPa)


Sand shear modulus (GPa)
Sand density (kg m3 )
Clay bulk modulus (GPa)
Clay shear modulus (GPa)
Clay density (kg m3 )
Reservoir porosity (per cent)
Clay content (per cent)

3643
3346
26402650
2034
719
23502680
1040
2050

2011 The Authors, GJI, 186, 825836


C 2011 RAS
Geophysical Journal International 

S-wave velocities and density to be calculated, which can in turn be


used in conjunction with the Zoeppritz equations to calculate the
P-wave reflection coefficient for a range of incident angles at the
reservoircaprock interface. By assuming prior uncertainty ranges
over the petrophysical properties in Tables 1 and 2 (which constitute
the model vector m in this case), prior parameter pdfs of the P-wave
velocity, S-wave velocity and density can be constructed. We assume
uniform pdfs over all parameter ranges in Tables 1 and 2.
Fig. 4 shows histograms of the models produced from 1000 000
random samples from the distributions in Tables 1 and 2 for each
of the three saturating fluids (gas, oil and water). Fig. 4 shows that
from the uniform parameter ranges in Tables 1 and 2 a large variety
of velocity and density models can be created.

3.2 Results
For all results crossplots have been discretized into 160 bins over
the range 0.5 to 0.3 in the intercept dimension and 200 bins over
the range 0.8 to 0.2 in the gradient dimension. For each survey
design the reservoir model has been sampled 200 000 times, and for
each particular reservoir model 50 realization of the data have been
produced by adding different realizations of data noise.

3.2.1 Porosity and saturating fluid


Guest & Curtis (2010) showed that the reservoir model parameter that has the largest effect on the survey design is the porosity.
Fig. 5(a) shows the information gain values as a function of P, the
zero offset receiver density for gas saturated reservoirs with low
(1020 per cent) and high (3040 per cent) Uniformly distributed
porosity ranges. The design corresponding to the maximum information gain is the optimal survey design. For a maximum offset of
30 and 300 receivers a P value of 10 receivers per degree equates to
a constant angular receiver separation; values of P less than 10 result
in more receivers being placed at larger offsets than near offsets,
and P values greater than 10 result in more receivers located at near
offsets than far offsets (Fig. 3). The results show a small increase
in the optimal zero offset receiver density as porosity increases: the
optimal zero offset receiver density (P) for the low porosity model
is seven receivers per degree whereas the optimal receiver density
for the high porosity model is nine receivers per degree.
Fig. 5(b) shows the information gain values as a function of P,
the zero offset receiver density, across the full prior porosity range
(Tables 1 and 2) for all three general reservoir models relating
to each of the three possible saturating fluids: oil, gas and brine.
Although the three reservoir models result in different information
gain values, the shape of the profiles are all similar with optimal
zero offset receiver densities of seven receivers per degree for the

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

The optimal survey design will be defined for a given prior Earth
model parameter probability distribution, (m). The model we use
is a simple two-layer reservoir (reservoir and caprock) which in
practical situations will be located under a possibly-complex overburden. We assume that rays have been traced through the overburden so that the angles of incidence of waves at the caprockreservoir
interface are known.
To evaluate the information measure in eq. (5) we need to define
prior probability distributions over i , i and i for i = 1, 2. We
assume the caprock is a shale with known = 3048 m s1 , =
1244 m s1 and = 2400 kg m3 , the same overburden model used
by Ostrander (1984) to analyse plane-wave reflection coefficients
as a function of incident angle for a reservoir model. The corresponding values of the lower layer (the reservoir) remain unknown.
In other scenarios the parameters of the upper layer or both layers
simultaneously could be assumed unknown, and the same methods
as below can be applied to calculate the optimal survey design. Here
we wished to study how the survey design depends on the reservoir
properties alone, so we held the caprock fixed.
The parameter vector m describes the reservoir rock properties.
A reservoir petrophysical model relates reservoir rock properties
to elastic and density parameters, and forms part of the forward
function F (m). We use the semi-empirical petrophysical model of
Goldberg & Gurevich (1998) which allows sand-shale reservoirs
with different percentage sand/shale ratios and different saturating
fluids to be analysed. For a particular set of rock physical properties
(Tables 1 and 2) this allows a corresponding set of P-wave and

829

830

T. Guest and A. Curtis

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

Figure 4. Velocity and density histograms for a gas-filled reservoir (a)(c), an oil-filled reservoir (d)(f) and a brine-filled reservoir (g)(i) using the parameter
values in Tables 1 and 2. Shading represents the histogram frequency.

oil and brine reservoirs and eight receivers per degree for the gas
saturated reservoir.
These results for the optimal designs are intuitive. For all of
the cases in Fig. 5 there is a larger proportion of receivers at far
offsets compared to near offsets. Since the data error increases with
offset, proportionally more receivers are required at large angles of

incidence to constrain the crossplot gradient compared with fewer


receivers required to constrain the reflection coefficient near zero
offset. The end-member survey designs only constrain either the
crossplot intercept (P = 20, Q = 0) or gradient (P = 0, Q = 20)
resulting in the other parameter having a high associated uncertainty.
However, although that much is intuitive, without performing the

C 2011 The Authors, GJI, 186, 825836
C 2011 RAS
Geophysical Journal International 

Design of industrial-scale 2-D seismic surveys

831

survey design algorithm the exact receiver density profile needed to


ensure optimality would remain unknown.
Although a difference in the optimal surveys as seen in each of
the plots in Fig. 5, the large difference in optimal designs observed
by Guest & Curtis (2010), particularly for different porosities, is not
apparent. This is because the range of incident angles considered in
Guest & Curtis (2010) extended to 70 whereas in this study they
never exceed the critical angle. Thus, we demonstrate that the nonlinearity in the forward function F around the critical angle creates
strong dependence of the optimal design on the particular reservoir
parameter ranges expected prior to conducting the survey. For precritical surveys on the other hand, there is (roughly speaking) a
one-size-fits-all design that has an optimal P value of around
eight when placing 300 receivers.

3.3 Number of receivers


In the above designs a total of 300 receivers have been placed.
However, for other cost or logistical constraints fewer (or more)
receivers may be required. In the following results the receiver
density values have been normalized so that P ranges from 0 to 1
from 1 to 0, with a value of 1 corresponding to the maximum
and Q
= 0.5 there is
receiver density possible in each case. When P = Q
a constant angular receiver separation.
Fig. 6 shows the normalized optimal receiver density designs
for the standard gas saturated reservoir for different porosity values
and total numbers of receivers located. The plot shows that when
fewer than about 250 receivers are used, the optimal survey depends
significantly on the number of receivers to be placed as shown by
the high P gradient in the horizontal direction. When placing fewer
than 250 receivers there is also a significant dependence on the
prior porosity range, particularly for higher porosity reservoirs. For
example, if only 100 receivers were to be placed, depending on the
prior porosity value, an optimal design could have a P value ranging
from less than 0.5 to 1.0.
When placing more than 250 receivers, the optimal P value is
always less than 0.5 representing an increasing receiver density with

C

2011 The Authors, GJI, 186, 825836


C 2011 RAS
Geophysical Journal International 

0.4

0.3

0.2

0.1

Figure 6. Optimal normalized P contours as a function of porosity and


total number of receivers. The 0.5 contour represents a constant receiver
separation with respect to angle. Values higher than 0.5 indicates a greater
receiver density at near offset angles and values less than 0.5 a greater
receiver density at far offsets.

offset, and the dependence on the prior porosity is significantly reduced. Fig. 6 also shows that the value of P never falls below 0.3 and
although not shown, we have checked that this result is also observed
when placing up to a total of 5000 receivers. This result implies that
once over a certain threshold of total-number-of-receivers-placed,
the relative distribution of receivers remains constant and defined
by P = 0.3 so that both the crossplot gradient and intercept can be
well constrained, even given the offset-dependent error (Fig. 1a).
The results in Fig. 6 imply that the one-size-fits-all design that
was evident for the porosity and fluid content does extend to all
porosity ranges and saturating fluids when the total number of receivers is greater than 250, but does not apply when the total number
of receivers is less than 250.
Although the results above show that the optimal designs for
the CMP gather using a linear receiver density distribution can
be expressed by a one-size-fits-all design once the total number
of receivers surpasses a threshold value, no measure has yet been
quantified of how much extra information about the subsurface

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

Figure 5. Information gain as a function of zero offset receiver density (P) for a survey consisting of 300 receivers. Plot (a) shows the results for a low porosity
(1020 per cent: solid line) and a high porosity (3040 per cent: dashed line) gas reservoir. Plot (b) shows the results for a gas (solid line), a brine (dashed line)
and an oil (dotted line) saturated reservoir for a reservoir with a uniformly distributed porosity range from 10 to 40 per cent.

832

T. Guest and A. Curtis

parameters is provided by the optimal design when compared to a


standard design of constant spatial receiver separation. Fig. 7 shows
the expected information gain values for an oil reservoir as a function of the total number of receivers placed when comparing the
optimal design with a standard design of equal spatial receiver separation for the single CMP gather. This plot shows the difference
between the information expected to be recorded (eq. 5) using a
standard survey design and that from an optimal design found using
the methods presented. The plot shows that for surveys consisting of
fewer than around 50 receivers, the optimal design provides significantly more information than a regularly spaced design. However,
the information gain provided by adding additional receivers to the
optimal design compared to simply performing a standard design
initially diminishes as the total number of receivers increases. This
agrees with the idea of diminishing returns which postulates that
as the number or receivers increases the relative advantage of using
optimal designs reduces (Coles & Morgan 2009).
The large change in expected information gain for low numbers
of receivers can also be explained by Fig. 6: as the number of receivers used increases from 0 to 200, the optimal design quickly
changes from one with the maximum number of receivers at small
offsets ( P = 1.0) towards a design which has equal angular receiver
spacing ( P = 0.5) which occurs at 180 receivers. Therefore, as the
number of receivers increases the optimal design tends to a design
of equal angular receiver spacing and as a result it is expected that
the relative information gain of the optimized survey will decrease.
However, as the number of receivers increases beyond 180 the P
value tends towards 0.3 corresponding to an optimal survey design
that becomes less like the standard design. Fig. 7 shows that this
results in an increased information gain with increasing number of
receivers, and although not shown we have tested that this remains
true up to 5000 receivers. This is in contrast to the idea of diminishing returns which would still expect a decrease in information
gain with increased number of receivers.
Although the methods used to locate the optimal receiver positions are different to those used by Guest & Curtis (2009) which
allowed receivers to be placed arbitrarily at any offset, the results
should be approximately consistent since both methods maximize
information about the same subsurface properties. We compare the
results of both methods when used to place 10 receivers (around

the maximum number able to be placed using the Guest & Curtis
(2009) method using a standard desktop PC (see Guest & Curtis
2010)) for a brine saturated reservoir within the angular range of
0 to 30 . For this comparison to be fair we have used a constant
error with offset, to be consistent with the results of Guest & Curtis
(2009).
Fig. 8 shows the cumulative number of receivers placed as a
function of incident angle for the Guest & Curtis (2009) method
(solid line), a P value of 1.0 (dashed line) equating to an optimal
survey when only 10 receivers are used, and a P value of 0.3 (dotted
line) which reflects the optimal design when more than 100 receivers
are placed. Fig. 8 shows that the results calculated using the Guest &
Curtis (2009) method in part match both results calculated using the
linear receiver density method. Optimal receivers are located at both
small offsets and large offsets to accurately estimate both the AVO
gradient and intercept with a region devoid of receivers between 10
and 22 offset. This is a result unobtainable in the examples above
due to our relatively coarsely parameterized design space (linearly
varying angular receiver density). As seen in Fig. 6 placing a low
number of receivers results in the optimal design being located in
a transition zone between a P value ranging between 0.3 and 1.0.
Fig. 8 shows that the method of Guest & Curtis (2009) spans both
of these. Since the Guest & Curtis (2009) method is restricted to
placing a maximum of around 10 receivers, it is impossible to say if
additional receivers would make the optimal result from that method
tend towards the optimal result of P = 0.3.
The above implies that a hyperparameterization using only two
hyperparameters is too coarse for the purpose of this comparison.
Fig. 9(a) shows example normalized receiver density profiles and
Fig. 9(b) the corresponding normalized cumulative receiver plots
is introduced
that become possible when a third hyperparameter ( M)
to represent the receiver density at 15 , half the maximum incident
angle, and when linear interpolation is used between 0 and 15 , and
between 15 and 30 . Adding an extra hyperparameter increases the
design space by one dimension but allows more variation in survey
designs.
Optimal surveys that use a total of 10, 20, 100 and 600 receivers
were calculated for a brine saturated reservoir (Tables 1 and 2) using
the three hyperparameter model. Fig. 10 shows how the three hyperparameter results (solid line) differ from the two hyperparameter

C 2011 The Authors, GJI, 186, 825836
C 2011 RAS
Geophysical Journal International 

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

Figure 7. Information gain expected from using the optimal receiver distribution compared to a standard survey design of equally-spaced receivers as
a function of the total number of receivers for a general oil filled reservoir.

Figure 8. Cumulative number of receivers placed for a brine filled reservoir


comparing the method of Guest & Curtis (2009) with the receiver density
method introduced above. The solid line represents the results found using
the Guest & Curtis (2009) method when 10 receivers are placed, the dashed
line the optimal receiver density result for a survey using 10 receivers, and
the dotted line an optimal survey when more than 100 receivers are placed.

Design of industrial-scale 2-D seismic surveys

833

Figure 9. (a) Normalized receiver density profiles for possible survey designs using three hyperparameters, and (b) the corresponding normalized cumulative
number of placed receivers, both as a function of incident angle. Note that in (a) the solid and dotted lines before and after 15 , respectively have been shifted
slightly so that they are visible.

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

Figure 10. Normalized cumulative placed receiver profiles for optimal surveys consisting of (a) 10 receivers, (b) 20 receivers, (c) 100 receivers and (d) 600
receivers. In each plot the dashed line represents the two hyperparameter result and the solid line the three hyperparameter result.

results (dashed line) for surveys consisting of (a) 10 receivers, (b)


20 receivers, (c) 100 receivers and (d) 600 receivers. When only 10
receivers are placed the two hyperparameter result has the highest
density of receivers at near offsets; with the three hyperparameter
result (Fig. 10a) the same result is seen but now all receivers are
located in the first 15 with no receivers between 15 and 30 . The
(the normalized receiver density at 15 ) for the 20, 100
value of M
and 600 receiver designs is 0 resulting in the inflection point seen
in Figs 10(b)(d). When using two hyperparameters the result for

C

2011 The Authors, GJI, 186, 825836


C 2011 RAS
Geophysical Journal International 

the 10 and 20 receiver designs are identical. However, when using three hyperparameters the results show a significant difference
with the 20 receiver design resembling the 100 and 600 receiver
designs. The addition of the extra hyperparameter now produces
optimal results that more closely resemble the result obtained using
the method of Guest & Curtis (2009). Fig. 11 shows that the 10
receiver results using the Guest & Curtis (2009) method are best
matched by the results found when placing 20 receivers using the
three hyperparameter method.

834

T. Guest and A. Curtis


The information gains calculated using the new designs compared
to a standard, equally spaced design result in values approximately
3 per cent higher than those seen in Fig. 7 for the three studied
models.

4 DISCUSSION

Figure 12. Normalized cumulative placed receivers for optimal threehyperparameter surveys consisting of 12 receivers (solid line), 13 receivers
(dot-dash line), 14 receivers (dashed line) and 15 receivers (dotted line).

Although it might initially seem worrying that the 10-receiver


results do not match using the two design methods, this is almost
certainly because the forward function F considered here differs
from that of Guest & Curtis (2009, 2010). The former studies assumed that the recorded amplitudes of arriving waves at each receiver would be inverted directly for petrophysical parameters using
the Zoeppritz equations and the petrophysical model of Goldberg
& Gurevich (1998). Here, however, we assume that recorded amplitudes will be summarized by AVO intercept and gradient parameters
as is standard practice in industry, and that these AVO parameters
will be inverted using eqs. (3). Hence, in each case the effective
data sets inverted differ, and so do the forward functions. Nevertheless, the similarity between the bold and dashed lines in Fig. 11
shows that the resulting designs in each case are strongly related, as
we would hope to be the case if the standard industrial AVO processing workflow is robust. We find that using the method herein,
the threshold at which the design shifts from that in Fig. 10(a) to
having an inflection point as in Fig. 10(b) occurs at 13 receivers
(Fig. 12 ).


C 2011 The Authors, GJI, 186, 825836
C 2011 RAS
Geophysical Journal International 

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

Figure 11. Cumulative number of receivers placed for a brine filled reservoir comparing the method of Guest & Curtis (2009) with the three hyperparameter design method. The thick solid line represents the results found
using the Guest & Curtis (2009) method when 10 receivers are placed, the
thin solid line represents the optimal receiver density result for a survey using 10 receivers, the dashed line an optimal survey when using 20 receivers,
the dot-dashed line an optimal survey when using 100 receivers and the
dotted line when 600 receivers are placed.

Although the designs showed that using three hyperparameters


seems to accurately represent the optimal design seen in Guest &
Curtis (2009), it is conceivable that adding a third hyperparameter
at 15 is not sufficient to allow all optimal designs found using the
iterative method of Guest & Curtis (2009) to be replicated. Using
a linear interpolation to calculate the receiver density between the
hyperparameters also restricts the range of possible designs. For
a given number of hyperparameters there therefore exists an optimization problem to locate the offsets at which the hyperparameters
are located, and to design the best method of receiver density interpolation between the hyperparameters. We do not address this
problem here as the comparison in Fig. 9 shows that the constraints
imposed by our choices of macro-parameterization do not seem to
restrict the range of possible designs unduly.
Nevertheless, as previously described, the 10 receiver, three hyperparameter case (represented by the thin solid line in Fig. 11)
does not match the result found by Guest & Curtis (2009) for 10
receivers (e.g. they have no receivers between 10 and 23 ). Using
four hyperparameter at incident angles 0 , 11 , 23 and 30 may
result in a closer match to the Guest & Curtis (2009) result and an
associated increase in information gain. Nevertheless, this gain is
likely to be small since the designs already have the freedom to be
fairly similar (shown by the similarity of the bold and dashed curves
in Fig. 11).
All of the results herein are based on the two-term Shuey approximation in eqs (2) and (3). Since we do observe some changes
in the optimal designs between the work of Guest & Curtis (2009)
and those found here due to the difference in the forward function employed, it is possible that different information gains would
be observed when using other AVO analysis methods [Connolly
(1999); Sheriff (1991); Whitcombe et al. (2002); Santos & Tygel
(2004); Morozov (2010)]. This remains to be tested.
Although the optimal surveys produced using our Bayesian design method result in information gains compared to standard constant spatial designs, the actual gain values are relatively small,
especially for large-scale industrial designs. In our analysis so far
we have not taken into consideration the extra cost factors (e.g. acquisition and processing costs) introduced when using an optimal
(non-regularly spaced) design, the cost of undertaking the design
process above, or the fact that surveys are generally designed to optimize noise attenuation and imaging and not solely to record data
for AVO processing. In practice additional cost factors should be
applied to Fig. 5 with a zero additional cost applied to the standard
design and a non-zero cost to all other designs with a magnitude
dependent on the extra costs expected to be incurred. In this way a
true optimal design could be determined.
Extra costs associated with using optimal designs are likely to be
significant. For marine seismics this would require that streamers
are redesigned with an extremely high associated cost. For conventional land seismics there would be significant extra expense due
to the need to survey and lay geophones over wide areas according to non-standard spatial templates. In both cases there would be
additional cost in adapting noise attenuation and imaging methods
to non-uniform receiver densities (in comparison to these, the costs

Design of industrial-scale 2-D seismic surveys

5 C O N C LU S I O N S
A Bayesian design method has been proposed which, when combined with a reservoir model and offset-dependent error measure,
produces industrial scale, optimal AVO designs that are shown to
decrease the expected uncertainty on the reservoir parameters compared to a standard design using the same number of receivers.
Although the optimal designs are similar for different porosity values and saturating fluids, the total number of receivers in the survey
has a large affect on the optimal design. However, once a particular
threshold on the total number of receivers has been passed there
exists a one-size-fits-all design that is optimal for any porosity,
fluid content or number of receivers.
Although these optimal designs provide extra information, the
CMP gather example analysed results in gains of up to only around
5 per cent when compared to a standard survey with constant spatial
receiver separation. Even when the reduced parameterisation is redefined to be more complex, these gains generally remain less than
around 10 per cent for surveys with more than 50 receivers. When
the cost of collecting and processing the new data is accounted for
it is unlikely that this increase in information will represent value
for money. For the given prior reservoir model and offset dependent

C

2011 The Authors, GJI, 186, 825836


C 2011 RAS
Geophysical Journal International 

error it is therefore concluded that although the one-size-fits-all


result shown above is optimal, when the cost of data collection and
processing are considered the current standard seismic survey design of constant spatial receiver separation is in fact optimal for
pre-critical AVO surveys. However, if the seismic system is one
in which the marginal cost is negative for switching off receivers
(such as a wireless data acquisition system in which data transmission costs dominate) the cost of data collection may actually be
reduced by using optimal designs to decide which sensors should
be transmitted and recorded.
AC K N OW L E D G M E N T S
Thanks are extended to Schlumberger for permission to publish this
work and to the Scottish Funding Council and the Edinburgh Collaborative of Subsurface Science and Engineering (ECOSSE) for
part funding this work. We would like to thank Emanuel Winterfors for his insight and stimulating discussions. Jeff Shragge and an
anonymous reviewer are thanked for their comments, which helped
to improve the manuscript.

REFERENCES
Ajo-Franklin, J., 2009. Optimal experiment design for time-lapse traveltime
tomography, Geophysics, 74(4), Q27Q40.
Atkinson, A. & Donev, A., 1992. Optimum Experimental Designs, Oxford
Science Publications, Oxford.
Barth, N. & Wunsch, C., 1990. Oceanographic experiment design by simulated annealing, J. Phys. Oceanogr., 20(9), 12491263.
Box, G. & Lucas, H., 1959. Design of experiments in nonlinear situations,
Biometrika, 46, 7790.
Carcione, J., Helle, H., Pham, N. & Toverud, T., 2003. Pore pressure estimation in reservoir rocks from seismic reflection data, Geophysics, 68(5),
15691579.
Castagna, J.P. & Swan, H.W., 1997. Principles of AVO crossplotting, Leading Edge, 16(4), 337344.
Chen, J. & Dickens, T.A., 2009. Effects of uncertainty in rock-physics
models on reservoir parameter estimation using seismic amplitude variation with angle and controlled-source electromagnetics data, Geophys.
Prospect., 57(1), 6174.
Chitwood, D., Tinnin, J., Hollis, C. & Hernandez, F., 2009. Cableless system
meets challenge of acquiring seismic to define subtle fractures in complex
shale, First Break, 27, 7985.
Clark, V., 1992. The effect of oil under in-situ conditions on the seismic
properties of rocks, Geophysics, 57(7), 894901.
Coles, D. & Curtis, A., 2010. A free lunch in azimuthally anisotropic survey
design, Comput. Geosci., doi:10.1016/j.cageo.2010.09.012, in press.
Coles, D. & Curtis, A., 2011. Efficient nonlinear Bayesian survey design
using DN optimization, Geophysics, 76(2), B1B5.
Coles, D. & Morgan, F., 2009. A method of fast, sequential experimental design for linearized geophysical inverse problems, Geophys. J. Int.,
178(1), 145158.
Connolly, P., 1999. Elastic impedance, Leading Edge, 18(4), 438452.
Curtis, A., 1999a. Optimal experiment design: cross-borehole tomographic
examples, Geophys. J. Int., 136, 637650.
Curtis, A., 1999b. Optimal design of focused experiments and surveys,
Geophys. J. Int., 139, 205215.
Curtis, A. & Lomax, A., 2001. Prior information, sampling distributions,
and the curse of dimensionality, Geophysics, 66(2), 372378.
Curtis, A. & Maurer, H., 2000. Optimizing the design of geophysical
experiments: is it worthwhile?, Leading Edge, 19(10), 10581062.
Curtis, A. & Wood, R., 2004. Optimal elicitation of probabilistic information
from experts, Geol. Soc. London Spec. Pub., 239, 127145.
Curtis, A., Michelini, A., Leslie, D. & Lomax, A., 2004. A deterministic
algorithm for experimental design applied to tomographic and microseismic monitoring surveys, Geophys. J. Int., 157, 595606.

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

of finding an optimal design are considered negligible). Thus, we


conclude that in practice, if we balance the magnitude of the gains
in information against the extra cost incurred for non-conventional
methods, the best surveys to use for AVO studies will in fact almost
always be regularly-spaced surveys. This is a somewhat surprising
result, given that standard surveys have been designed to simplify
and aid noise attenuation and imaging rather than for petrophysical inversion. However, it does explain why these standard designs
have also been used so successfully for petrophysical inversion in
the past. The relative drop in information resulting from designing
for noise attenuation and imaging rather than for AVO is generally
lower than 10 per cent.
The conclusion reached that standard designs are optimal is
mainly because altering streamer and cable designs has a high associated, positive cost function. However, we note that wireless land
seismic acquisition methods, such as the FireFly system, are becoming increasingly popular. In such a system, single receivers are
wirelessly connected to a central recording facility without the need
for cables, thus removing the large cost penalty of using optimal
designs over standard designs (Chitwood et al. 2009). A typical
survey using the FireFly system can consist of over 10 500 receivers and 7000 shot points. The main cost is then associated with
data transmission, sorting and storage of the data. Hence, switching off unnecessary receivers for each shot has a negative associated cost. The methods described above can therefore be used to
generate optimal recording-receiver designs so as to maximally
record subsurface information whilst also reducing acquisition
costs.
Since the algorithm calculates the optimal incident angles at
the caprock/reservoir boundary it is easy to transform the results
into specific, more complex geometrical cases. Although the design may change if layers dip instead of being horizontal, the algorithm that we use to calculate optimal designs would still be
robust. This is because we calculate the optimal distribution of
incident angles at the reflector to be analysed. Whatever distribution of angles are found, these can be traced back to the surface
through the overburden model to find optimal sensor locations on the
surface.

835

836

T. Guest and A. Curtis


Modern Approaches in Geophysics, Vol. 18, ch. 3, eds Thurber, C. &
Rabinowitz, N., Kluwer Academic Publishers, London.
Santos, L. & Tygel, M., 2004. Impedance-type approximations of the PP
elastic reflection coefficient: modeling and AVO inversion, Geophysics,
69(2), 592598.
Shannon, C., 1948. A mathematical theory of communication, Bell System
Tech. J., 27, 623656.
Sheriff, R., 1991. Encyclopedic Dictionary of Exploration Geophysics, SEG.
Shewry, M.C. & Wynn, H.P., 1987. Maximum entropy sampling, J. Appl.
Stat., 14, 165170.
Shuey, R.T., 1985. A simplification of the Zoeppritz equations, Geophysics,
50(4), 609614.
Steinberg, D., Rabinowitz, N., Shimshoni, Y. & Mizrachi, D., 1995. Configuring a seismographic network for optimal monitoring of fault lines and
multiple sources, Bull. seism. Soc. Am., 85(6), 18471857.
Stummer, P., Maurer, H. & Green, A.G., 2004. Experimental design: electrical resistivity data sets that provide optimum subsurface information,
Geophysics, 69(1), 120139.
Tarantola, A., 2005. Inverse Problem Theory and Methods for Model Parameter Estimation, SIAM.
Van den Berg, J., Curtis, A. & Trampert, J., 2003. Optimal nonlinear bayesian
experimental design: an application to amplitude versus offset experiments, Geophys. J. Int., 155, 411421.
Van den Berg, J., Curtis, A. & Trampert, J., 2005. Corrigendum, Geophys.
J. Int., 161, 265265.
Whitcombe, D., Connolly, P., Reagan, R. & Redshaw, T., 2002. Extended
elastic impedance for fluid and lithology prediction, Geophysics, 67(1),
6367.
Wilkinson, P., Meldrum, P., Chambers, J., Kuras, O. & Ogilvy, R., 2006. Improved strategies for the automatic selection of optimized sets of electrical
resistivity tomography measurement configurations, Geophys. J. Int., 167,
11191126.
Winterfors, E. & Curtis, A., 2008. Numerical detection and reduction
of non-uniqueness in nonlinear inverse problems, Inverse Probl., 24(2),
025016, doi:10.1088/0266-5611/24/2/025016.
Winterfors, E. & Curtis, A., 2010. A bifocal measure of expected ambiguity
in bayesian nonlinear parameter estimation, Technometrics,, submitted.
Zoeppritz, K., 1919. On the reflection and propagation of seismic waves,
Gottinger Nachrichten, 1, 6684.


C 2011 The Authors, GJI, 186, 825836
C 2011 RAS
Geophysical Journal International 

Downloaded from http://gji.oxfordjournals.org/ by guest on October 24, 2015

Furman, A., Ferre, T.P.A. & Warrick, A.W., 2004. Optimization of ert
surveys for monitoring transient hydrological events using perturbation sensitivity and genetic algorithms, Vadose Zone J., 3(4), 1230
1239.
Furman, A., Ferre, T. & Heath, G., 2007. Spatial focusing of electrical resistivity surveys considering geologic and hydrologic layering, Geophysics,
72(2), F65F73.
Goldberg, I. & Gurevich, B., 1998. A semi-empirical velocity-porosity-clay
model for petrophysical interpretation of p- and s-velocities, Geophys.
Prospect., 46, 271285.
Guest, T. & Curtis, A., 2009. Iteratively constructive sequential design of
experiments and surveys with nonlinear parameter-data relationships, J.
geophys. Res., 114, B04307, doi:10.1029/2008JB005948.
Guest, T. & Curtis, A., 2010. Optimal trace selection for AVA processing
of shale-sand reservoirs, Geophysics, 75(4), C37C47.
Kijko, A., 1977a. An algorithm for the optimal distribution of a regional
seismic network - 1, Pure appl. Geophys., 115(4), 9991009.
Kijko, A., 1977b. An algorithm for the optimum distribution of a regional
seismic network - 2. an analysis of the accuracy of location of local
earthquakes depending on the number of seismic stations, Pure appl.
Geophys., 115(4), 10111021.
Marion, D., Nur, A., Yin, H. & Han, D., 1992. Compressional velocity and
porosity in sand-clay mixture, Geophysics, 57(4), 554563.
Maurer, H. & Boerner, D., 1998a. Optimized design of geophysical experiments, Leading Edge, 17(8), 11191125.
Maurer, H. & Boerner, D., 1998b. Optimized and robust experimental
design: a non-linear application to em sounding, Geophys. J. Int., 132,
458468.
Maurer, H., Boerner, D. & Curtis, A., 2000. Design strategies for electromagnetic geophysical surveys, Inverse Probl., 16, 10971117.
Mavko, G., Mukerji, T. & Dvorkin, J., 1998. The Rock Physics Handbook,
Cambridge University Press, Cambridge.
Morozov, I., 2010. Exact elastic P/SV impedance, Geophysics, 75(2),
C7C13.
Ostrander, W.,
1984. Plane-wave reflection coefficients for gas
sands at nonnormal angles of incidence, Geophysics, 49(10), 1637
1648.
Rabinowitz, N. & Steinberg, D., 2000. A statistical outlook on the problem
of seismic network configuration, in Advances in Seismic Event Location,

You might also like